首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 629 毫秒
1.
Deuterium nuclear magnetic resonance spectroscopy (2H NMR) has been employed to investigate the interaction of lung type II myosin protein with neutral bilayers containing dimyristoylphosphatidylcholine (DMPC) as the only constituent and mixed bilayers containing the negatively charged lipid dimyristoylphosphatidylglycerol (DMPG). DMPC was deuterated at its headgroup by substituting the four protons at the alpha- and beta-positions (DMPC-d4) and the nine protons at the gamma-position (DMPC-d9). DMPG was perdeuterated at its headgroup (DMPG-d5). No changes were observed in the quadrupole splittings or spin-lattice relaxation times for the deuterated DMPC headgroup segments when increasing amounts of myosin were added to liposomes, made exclusively of DMPC-d9 or of DMPC-d4. However, upon the insertion of the negatively charged lipid DMPG at 1:1 molar ratio into the DMPC bilayers, myosin was found to interact electrostatically with the liposomes, thereby affecting significantly both the quadrupole splittings and spin-lattice relaxation rates of the alpha-, beta-, and gamma-deuterons in labeled DMPC. Monitoring DMPG-d5 in mixed DMPC/DMPG bilayers revealed a direct electrostatic interaction of DMPG with the protein, where positively charged lysine residues located at the tail domain of myosin provide the necessary sites for the interaction to occur. When ATP and Mg2+ were complexed to the head domain of myosin, a reduced interaction with the negatively charged bilayers was observed. The results clearly indicate that a type II myosin can interact with membranes without the need for a specific hydrophobic domain or an anchor in the protein molecule, provided that negatively charged lipids are present in the bilayer.  相似文献   

2.
The addition of lanthanides (Tm3+, Yb3+, Er3+, or Eu3+) to a solution of long-chain phospholipids such as dimyristoylphosphatidylcholine (DMPC) and short-chain phospholipids such as dihexanoylphosphatidylcholine (DHPC) is known to result in a bilayer phase in which the average bilayer normal aligns parallel to an applied magnetic field. Lanthanide-doped bilayers have enormous potential for the study of membrane proteins by solid-state NMR, low-angle diffraction, and a variety of optical spectroscopic techniques. However, the addition of lanthanides poses certain challenges to the NMR spectroscopist: coexistence of an isotropic phase and hysteresis effects, direct binding of the paramagnetic ion to the peptide or protein of interest, and severe paramagnetic shifts and line broadening. Lower water concentrations and larger DMPC/DHPC ratios than those typically used in bicelles consistently yield a single oriented bilayer phase that is stable over a wide range of temperature (approximately 35-90 degrees C). Among the above choice of lanthanides, Yb3+ is found to give minimal paramagnetic shifts and line broadening at acceptably low concentrations necessary for alignment (i.e., Yb3+/DMPC mole ratios equal to or greater than 0.01). Finally, the addition of a phospholipid chelate, 1,2-dimyristoyl-sn-glycero-3-phosphoethanolamine--diethylenetriaminepent aacetic acid, is observed to significantly reduce paramagnetic broadening and presumably prevent direct association of the peptide with the lanthanide ions.  相似文献   

3.
Lipid activation of protein kinase C alpha (PKC alpha) was studied by using a model mixture containing 1, 2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC), 1, 2-dimyristoyl-sn-glycero-3-phosphoserine (DMPS), and 1, 2-dimyristoyl-sn-glycerol (1,2-DMG). This lipid mixture was physically characterized by differential scanning calorimetry (DSC), Fourier transform infrared spectroscopy (FTIR), and 31P-nuclear magnetic resonance (31P-NMR). Based on these techniques, a phase diagram was constructed by keeping a constant DMPC/DMPS molar ratio of 4:1 and changing the concentration of 1,2-DMG. This phase diagram displayed three regions and two compounds: compound 1 (C1), with 45 mol% 1,2-DMG, and compound 2 (C2), with 60 mol% 1,2-DMG. When the phase diagram was elaborated in the presence of Ca2+ and Mg2+, at concentrations similar to those used in the PKC alpha activity assay, the boundaries between the regions changed slightly and C1 had 35 mol% 1,2-DMG. The activity of PKC alpha was studied at several temperatures and at different concentrations of 1,2-DMG, with a maximum of activity reached at 30 mol% 1,2-DMG and lower values at higher concentrations. In the presence of Ca2+ and Mg2+, maximum PKC alpha activity occurred at concentrations of 1,2-DMG that were close to the boundary in the phase diagram between region 1, where compound C1 and the pure phospholipid coexisted in the gel phase, and region 2, where compounds C1 and C2 coexisted. These results suggest that the membrane structure corresponding to a mixture of 1,2-DMG/phospholipid complex and free phospholipid is better able to support the activity of PKC alpha than the 1,2-DMG/phospholipid complex alone.  相似文献   

4.
Turbidity (absorbance at 470 nm) measurements revealed human serum low density lipoprotein (LDL) to cause, within a few minutes and at physiological pH and [NaCl], the aggregation of liquid crystalline large unilamellar liposomes (LUVs) of dimyristoylphosphatidylglycerol (DMPG). No evidence for concomitant lipid or aqueous contents mixing was obtained with fluorescent assays for these processes, in keeping with the lack of fusion of LUVs. Involvement of apoB is implicated by the finding that tryptic digestion of LDL abrogates its ability to cause aggregation. Aggregation is not caused by VLDL, HDL2, or HDL3. Interestingly, also oxidised LDL failed to aggregate DMPG vesicles. Aggregation of DMPG LUVs by LDL did depend on the ionic strength of the medium as well as on the phase state of the lipid. More specifically, below the main transition temperature Tm maximal aggregation was seen in the presence of 25-100 mM NaCl, whereas slightly higher (up to 150 mM) [NaCl] were required when T>Tm. Aggregation due to LDL was also observed for dimyristoylphosphatidylserine as well as for dipalmitoylphosphatidylglycerol LUVs, whereas liposomes composed of either unsaturated acidic phospholipids or different phosphatidylcholines were not aggregated. Involvement of electrostatic attraction between the acidic phosphate of DMPG and cationic residues in apoB is suggested by the finding that increasing the content of dimyristoylphosphatidylcholine (DMPC) in DMPG liposomes reduced their aggregation and at XDMPC=0.50 no response was evident. Notably, increasing the mole fraction of 1-palmitoyl-2-oleyl-PG (POPG) in DMPG LUVs progressively reduced their aggregation by LDL and at XPOPG=0.50 there was complete inhibition. The latter effect of POPG is likely to be due to augmented hydration of the unsaturated lipid constituting a barrier for the contact between apoB and the vesicle surface. In keeping with this view, the presence of the strongly hygroscopic polymer, poly(ethylene glycol) at 1% (by weight) enhanced the aggregation and could partly reverse the inhibition by POPG.  相似文献   

5.
Deuterium NMR spectroscopy has been employed to characterize the concentration dependence of orientational order in DMPC/DHPC bicellar solutions with molar ratios q = [DMPC]/[DHPC] = 3.3, 2.7, and 2.3. The stability of a discotic nematic phase can, in general, be predicted from a simple Onsager picture involving the size and concentration of the mesogenic unit, but for the bicellar solutions this model is not adequate. Specifically, macroscopic alignment is observed at total lipid concentrations well below that, 1-10% (w/w) predicted by Onsager's model. Thus the discotic nematic phase is stable to approximately 3-5% (w/w) for q = 3.3-2.3, and the bicellar order is highest just before phase separation occurs at the minimum total phospholipid concentration. This implies the presence of a DHPCbic -->/<-- DHPCsol equilibrium in establishing bicellar size, thereby extending the range of concentrations for which alignment occurs. Bicellar morphology has been verified for a wide range of concentrations, temperatures, and q-values, but as viscosity measurements demonstrate, major morphological changes take place as the temperature is reduced below 30 degreesC.  相似文献   

6.
1. The phosphonium analogues of choline, phosphorylcholine, CDPcholine and phosphatidylcholine were synthesized chemically and characterized by 1H-NMR and 31P-NMR; in 1,2-distearoyl-DL-glycero-3-phosphorylphosphocholine, the 31P-NMR chemical shift of phosphonium relative to phosphate was--28.2 ppm. 2. A comparison was made of the rates of reaction of choline kinase, cholinephosphate cytidyltransferase, cholinephosphotransferase and phospholipase C on natural and phosphonium substrates. Enzyme reaction rates were similar for all but the cytidyltransferase, which exhibited a 3-fold preference for the normal substrate. 3. Weanling rats were maintained for 6 weeks on a diet in which choline was fully replaced by phospho[1,2-14C2]choline mixed with a trace of [Me-3H] choline. Incorporation of phosphocholine into liver lipids was detectable by 31P-NMR even in crude tissue homogenates. Choline-based phospholipids of liver, kidney, lung and brain were extracted, and phosphocholine incorporation calculated from 31P-NMR peak area ratios. The phosphatidylcholine analogues were separated by preparative thin-layer chromatography. Incorporation of phosphocholine ranged from 33% in lung phosphatidylcholine to 6% in kidney sphingomyelin. Variations in 14C/3H ratio between feed and phospholipid extracts indicated preferences for exogenous choline over phosphocholine varying from 1.3: 1 in brain to 3.2: 1 in liver. The results indicated that phosphocholine is a potentially useful 31P-NMR probe for the study of membrane lipids.  相似文献   

7.
Perdeuterated indole-d6 and N-methylated indole-d6 were solubilized in lamellar liquid crystalline phases composed of either 1,2-diacyl-glycero-3-phosphocholine (14:0)/water or 1,2-dialkyl-glycero-3-phosphocholine(14:0/water. The molecular ordering of the tryptophan analogs was determined from deuteron quadrupole splittings observed in 2H-NMR spectra on macroscopically aligned lipid bilayers. NMR spectra were recorded with the bilayers oriented perpendicular to or parallel with the external magnetic field, and the values of the splittings differed by a factor of 2 between these distinct orientations, indicating fast rotational motion of the molecules about an axis parallel to the bilayer normal. In all cases the splittings were found to decrease with increasing temperature. Relatively large splittings were observed in all systems, demonstrating that the tryptophans partition into a highly anisotropic environment. Solubilization most likely occurs at the lipid/water interface, as indicated by 1H-NMR chemical shift studies. The 2H-NMR spectra obtained for each analog were found to be rather similar in ester and ether lipids, but with smaller splittings in the ether lipid under similar conditions. The difference was slightly less for the indole molecule. Furthermore, in both lipid systems the positions of the splittings from indole were different from those of N-methyl indole. The results suggest that 1) the tryptophan analogs are solubilized in the interfacial region of the lipid bilayer, 2) the behavior may be modulated by hydrogen bonding in the case of indole, and 3) hydrogen bonding with the lipid carbonyl groups is not likely to play a major role in the solubilization of single indole molecules in the ester lipid bilayer interface.  相似文献   

8.
Cadmium(II) interactions with multilamellar vesicles of dimyristoyl (DM)- and dipalmitoyl (DP)-phosphatidylcholine (PC), -phosphatidylserine (PS), -phosphatidic acid (PA), -phosphatidylglycerol (PG) and -phosphatidylethanolamine (PE) have been investigated both from the metal and the membrane viewpoints, respectively, by solution 113Cd-NMR and diphenylhexatriene fluorescence polarization coupled with solid-state 31P-NMR. Results can be summarized as follows. (1) Strong cadmium binding to membrane phospholipids results in a decrease of the free Cd(II) 113Cd-NMR isotropic signal and because of slow exchange, in the NMR time scale, between free and bound cadmium pools, the lipid/water partition coefficients, Klw, of the Cd(II) species can be determined in the lamellar gel (fluid) phase. It is found Klw(DMPC) approximately Klw(EggPE) approximately 2+/-2 (2+/-2); Klw(DMPA)=392+/-20 (505+/-25); Klw(DMPG)=428+/-21 (352+/-17); Klw(DMPS)=544+/-27 (672+/-34). Cadmium interactions with membrane phospholipids are therefore electrostatic in nature and the phosphate moiety is proposed as a potential binding site. (2) The presence of Cd(II) stabilizes the gel phases of PG, PA and PS lipids and leads to suppression of the main phase transition for PA and PS. These effects are reduced upon increasing salinity to 0.5 M Cl- and abolished at 1.8 M Cl-, Cd(II) being removed from the membranes due to formation of soluble CdCln species. Moving the pH from 7 to 6 also decreases Cd(II) binding to PA, because of surface charge reduction. (3) Cadmium promotes the formation of isotropic 31P-NMR lines with PG systems and of a hexagonal phase on egg PE bilayers at 24 degreesC, suggesting dramatic membrane reorganization. Properties of cadmium and calcium interacting with phospholipid model membranes are compared, and the potential roles of these interactions in the molecular mechanisms of cadmium uptake and toxicity in cells are discussed.  相似文献   

9.
A stable smectic phospholipid bilayer phase aligned with the director parallel to the magnetic field can be generated by the addition of certain trivalent paramagnetic lanthanide ions to a bicellar solution of dimyristoylphosphatidylcholine (DMPC) and dihexanoylphosphatidylcholine (DHPC) in water. Suitable lanthanide ions are those with positive anisotropy of their magnetic susceptibility, namely Eu3+, Er3+, Tm3+, and Yb3+. For samples doped with Tm3+, this phase extends over a wide range of Tm3+ concentrations (6-40 mM) and temperatures (35-90 degrees C) and appears to undergo a transition from a fluid nematic discotic to a fluid, but highly ordered, smectic phase at a temperature that depends on the thulium concentration. As a membrane mimetic, these new, positively ordered phospholipid phases have high potential for structural studies using a variety of techniques such as magnetic resonance (EMR and NMR), small-angle x-ray and neutron diffraction, as well as optical and infrared spectroscopy.  相似文献   

10.
Metabolic factors that influence the transition form reversible to irreversible ischemic injury were studied in the rat liver in vivo with 31P-nuclear magnetic resonance (NMR) spectroscopy. Hepatic ischemia for 15, 35, or 65 min was produced by occlusion of the hepatic artery and portal vein in rats. Ischemia caused a rapid decrease in the ATP concentration ([ATP])-to-P(i) concentration ratio and pH within 5 min, but there was little change in these variables detectable by 31P-NMR with longer periods of ischemia. After reperfusion, the [ATP] and P(i) concentration returned toward normal values in livers exposed to 15 or 35 min of ischemia, but 65 min of ischemia were associated with only modest recovery in [ATP], and the [ATP] later decreased. Because the 31P-NMR spectrum was similar after brief compared with prolonged ischemia, it appears that neither ATP depletion, P(i) accumulation, nor acidosis predicts metabolic recovery. Hepatic intracellular NA+ was also measured in separate groups of animals by 23Na-NMR in the presence of a shift agent, thulium (III) 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetrakis (methylene-phosphonate) (TmDOTP5-), and by atomic absorption spectroscopy. Under baseline conditions, the concentration of intracellular Na+ was 15.2 mM by atomic absorption spectroscopy and 16.5 mM by 23Na-NMR. Although the 31P-NMR spectrum responded very rapidly to the onset of ischemia, intracellular Na+ concentration measured by 23Na-NMR increased gradually but steadily at approximately 1.0 mM/min during early (up to 15 min) ischemia. These observations demonstrate that a rise in intracellular Na+ does occur early ischemia, that TmDOTP5- can be applied in vivo for analysis of intracellular Na+ in the ischemic liver, and that 31P-NMR spectroscopy is very sensitive to early ischemic injury.  相似文献   

11.
Tetracaine-induced biphasic structure-functional alterations were investigated in acetylcholinesterase (AChE)-associated membrane vesicles from the electric organ of Torpedo californica. Enzyme assays showed that tetracaine exhibits a biphasic effect on the activity of membrane-bound AChE: increasing it at low concentrations (< 12 mM) and decreasing it at high concentrations (> 12 mM). Fluorescence-polarization experiments demonstrated that tetracaine affects the fluidity of lipid hydrocarbon chains of these membranes in a biphasic manner: increasing it at < 20 mM and decreasing it at > 20 mM. This small molecule also alters the fluidity of the negatively charged lipid head group: increasing it at < 13 mM and remaining essentially at the same level at > 13 mM. The positively charged lipid head group is unaffected. Contrasting effects on AChE activity with changes in membrane fluidity showed that [tetracaine] for AChE activity is comparable to that for the fluidity of the negatively charged lipid head group (12 mM versus 13 mM), but lower than that for a biphasic effect on the fluidity of lipid hydrocarbon chains (12 mM versus 20 mM). Differential scanning microcalorimetry showed that, due to membrane protein-lipid interaction, the lipid-phase transition temperature (tml) is higher for AChE-associated membrane vesicles than for isolated lipids from these membranes. An overall disordering of the membranes by tetracaine, as inferred from the lowering of tml, was also demonstrated. These findings suggested that binding of tetracaine to the lipid polar head group and membrane protein-lipid interaction may contribute to a higher [tetracaine] in inducing a comparable biphasic effect on membrane fluidity. At high [tetracaine], charge interactions between the tetracaine cation and the negatively charged lipid head group may result in a new lipid phase in the membranes, which could reverse the increase in membrane fluidity, resulting in the observed biphasic effect. Although both tetracaine and alcohol are amphiphilic species, they exhibit distinctive structure-functional effects on the membranes, as shown by comparing the results obtained on tetracaine with those previously reported for alcohol. The present observations may have significant physiological implications and may be of importance in understanding the biochemical effects of tetracaine in correlation with its physiological impact.  相似文献   

12.
Polymer-free and polymer-bearing liposomes from dimyristoylphosphatidylcholine (DMPC), dimyristoylphosphatidylglycerol (DMPG) and cholesterol (Ch) 10:1:1 (w/w/w) were prepared. Polymer-bearing liposomes were formed by incorporating an uncharged polymer [poly(vinyl alcohol) (PVA), poly(vinyl alcohol-co-vinylacetal) (PVA-Al), poly(vinyl alcohol-co-vinyl propional) (PVA-Prol) poly(vinyl alcohol-co-vinyl butiral) (PVA-Bul) copolymer, polyvinyl pyrrolidone (PVP) or polyethylene glycol (PEG) respectively]. The effect of some polymers on the phase transition parameter of phospholipids (DMPC and DMPG) has been studied by differential scanning calorimetry (DSC). DSC has become a standard technique for studying the thermally induced transition of phospholipid molecules from an ordered crystalline-like state at low temperature (gel phase) to a liquid crystalline-like state at higher temperature. The interaction between phospholipids and polymers was characterized by the alteration of the phase transition parameters (Tp, Tm, delta T1/2, delta Hp, delta Hm) measured by DSC. It was observed that some phase transition parameters were affected by all polymers but in different extent. PVA-Prol copolymer exhibited a prominent effect in increasing the membrane cooperativity. The influence of PVA-Bul was unique; it decreased the cooperativity of phospholipid molecules. It was also shown that the lipid membranepolymer interaction considerably depends on both the chemical structure and the molecular mass of the polymers.  相似文献   

13.
The effect of pH and cholesterol on the dimyristoylphosphatidic acid (DMPA) model membrane system has been investigated by solid state 2H- and 31P-NMR. It has been shown that each of the three protonation states of the DMPA molecule corresponds to a 31P-NMR powder pattern with characteristic delta sigma values; this implies additionally that the proton exchange on the membrane surface is slow on the NMR time scale (millisecond range). Under these conditions, the 2H-labeled lipid chains sense only one magnetic environment, indicating that the three spectra detected by 31P-NMR are related to charge-dependent local dynamics or orientations of the phosphate headgroup or both. Chain ordering in the fluid phase is also found to depend weakly on the charge at the interface. In addition, it has also been found that the first pK of the DMPA membrane is modified by changes in the lipid lateral packing (gel or fluid phases or in the presence of cholesterol) in contrast to the second pK. The incorporation of 30 mol% cholesterol affects the phosphatidic acid bilayer in a way similar to what has been reported for phosphatidylcholine/cholesterol membranes, but to an extent comparable to 10-20 mol % sterol in phosphatidylcholines. However, the orientation and molecular order parameter of cholesterol in DMPA are similar to those found in dimyristoylphosphatidylcholine.  相似文献   

14.
Mismatch between the hydrophobic thicknesses of transmembrane proteins and the supporting lipid bilayer and its consequences on the lateral organization of lipids have been investigated with bacteriorhodopsin and phosphatidylcholine species with a variety of acyl-chain lengths. The purple membrane, from the bacterium Halobacterium halobium, was used and reconstituted with dilauroyl-(Lau2GroPCho), dimyristoyl- (Myr2GroPCho), dipalmitoyl- (Pam2GroPCho) and distearoyl- (Ste2GroPCho) glycerophosphocholine. The phase behaviour of the lipids was investigated at different temperatures and different protein/lipid molar ratios, by analyzing the fluorescence excitation spectra of the 1-acyl-2-[8-(2-anthroyl)-octanoyl]-sn-glycero-3-phosphocholine probe, and by measuring the fluorescence depolarization of the 1,6-diphenyl-1,3,5-hexatriene probe. Data obtained with 1-acyl-2-[8-(2-anthroyl)-octanoyl]-sn-glycero-3-phosphocholine shows that bacteriorhodopsin produced positive or negative shifts in the phase transition temperature of the host lipids depending on the strength and sign of the mismatch between the lipid and protein hydrophobic thicknesses and also on the protein concentration and aggregation state in the lipid bilayer. In the region of high protein concentration (bacteriorhodopsin/phosphatidylcholine molar ratios approximately 1:50) and despite the presence of the endogenous lipids, bacteriorhodopsin (hydrophobic length dP approximately 3.0-3.1 nm) brought about a large upward shift in the phase-transition temperature of Lau2GroPCho (delta T approximately 40 K, mean hydrophobic thickness d approximately 2.4 nm), and to a lesser extent of Myr2GroPCho (delta T approximately 23 K, d approximately 2.8 nm), accounting for a strong rigidifying effect of the protein on these short-chain lipids. Bacteriorhodopsin had no influence on the phase properties of Pam2GroPCho (delta T approximately 0 K, d approximately 3.2 nm), a lipid whose mean hydrophobic thickness is similar to that of the protein. In contrast, the transition temperature of Ste2GroPCho was decreased (delta T approximately -13 K, d approximately 3.7 nm), indicating a fluidifying effect of the protein on this long-chain lipid. Similar effects on the lipid acyl-chain order were observed in the region of high-protein dilution (bacteriorhodopsin/phosphatidylcholine molar ratios < 1:500). In this region and for Lau2GroPCho, both the spectroscopic data and circular-dichroism spectra indicated that the protein was in the monomeric form. Phase diagrams, in temperature versus bacteriorhodopsin concentration, were constructed for Lau2GroPCho and Ste2GroPCho. On account of microscopic theoretical models and of the relative values of dP and d, these diagrams indicate a preference of the protein for those lipid molecules which are in the gel-ordered state in Lau2GroPCho but in the liquid disordered state in Ste2GroPCho.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

15.
The stabilities of two types of polymerically stabilized liposomes consisting of PEG-grafted (DSPC:CHOL:DSPE-PEG1900, 5:4:1) and physically adsorbed carboxymethylchitin (CMC) and carboxymethyl/glycolchitin (CO) are compared. The polyelectrolyte is adsorbed on positive (DSPC:CHOL:DMTAP, 5:4:1) and neutral (DSPC:CHOL, 1:1) liposomes at different molecular weights (Mw). In PBS buffer (c(s) = 154 mM, pH = 7.4) the theoretical stability ratios (W) calculated using the classical DLVO Theory, indicate that the CMC-coated vesicles and the negative liposomes (DSPC:CHOL:DMPG, 5:4:1) are highly stable (W > 1) compared to the PEG-grafted (W = 0.9511) and CO-coated (W = 0.9550) liposomes. Meanwhile, experimentally determined values of W, prove that the PEG-grafted is the most stable suspension (W = 5.5). Computation of the theoretical values of W for liposome-red blood cell and liposome-macrophage indicates that the electrosterically stabilized suspensions and the negative liposomes are stable. Light scattering results show that the flocculation of liposomes in blood and plasma depends on polymer molecular weight, type of polyelectrolyte and surface charge of the uncoated liposome. Neutral liposomes coated with CMC of Mw = 1.01 x 10(5) and negative liposomes provide a more effective barrier to plasma macromolecular protein adsorption than the grafted PEG groups and are easy to resuspend in blood.  相似文献   

16.
In the present study we show that K+/H+ hydroxyl-containing ionophores lasalocid-A (LAS) and nigericin (NIG) in the nanomolar concentration range, inhibit Fe2+-citrate and 2,2'-azobis(2-amidinopropane) dihydrochloride (ABAP)-induced lipid peroxidation in intact rat liver mitochondria and in egg phosphatidylcholine (PC) liposomes containing negatively charged lipids--dicetyl phosphate (DCP) or cardiolipin (CL)--and KCl as the osmotic support. In addition, monensin (MON), a hydroxyl-containing ionophore with higher affinity for Na+ than for K+, promotes a similar effect when NaCl is the osmotic support. The protective effect of the ionophores is not observed when the osmolyte is sucrose. Lipid peroxidation was evidenced by mitochondrial swelling, antimycin A-insensitive O2 consumption, formation of thiobarbituric acid-reactive substances (TBARS), conjugated dienes, and electron paramagnetic resonance (EPR) spectra of an incorporated lipid spin probe. A time-dependent decay of spin label EPR signal is observed as a consequence of lipid peroxidation induced by both inductor systems in liposomes. Nitroxide destruction is inhibited by butylated hydroxytoluene, a known antioxidant, and by the hydroxyl-containing ionophores. In contrast, valinomycin (VAL), which does not possess alcoholic groups, does not display this protective effect. Effective order parameters (Seff), determined from the spectra of an incorporated spin label are larger in the presence of salt and display a small increase upon addition of the ionophores, as a result of the increase of counter ion concentration at the negatively charged bilayer surface. This condition leads to increased formation of the ion-ionophore complex, the membrane binding (uncharged) species. The membrane-incorporated complex is the active species in the lipid peroxidation inhibiting process. Studies in aqueous solution (in the absence of membranes) showed that NIG and LAS, but not VAL, decrease the Fe2+-citrate-induced production of radicals derived from piperazine-based buffers, demonstrating their property as radical scavengers. Both Fe2+-citrate and ABAP promote a much more pronounced decrease of LAS fluorescence in PC/CL liposomes than in dimyristoyl phosphatidylcholine (DMPC, saturated phospholipid)-DCP liposomes, indicating that the ionophore also scavenges lipid peroxyl radicals. A slow decrease of fluorescence is observed in the latter system, for all lipid compositions in sucrose medium, and in the absence of membranes, indicating that the primary radicals stemming from both inductors also attack the ionophore. Altogether, the data lead to the conclusion that the membrane-incorporated cation complexes of NIG, LAS and MON inhibit lipid peroxidation by blocking initiation and propagation reactions in the lipid phase via a free radical scavenging mechanism, very likely due to the presence of alcoholic hydroxyl groups in all three molecules and to the attack of the aromatic moiety of LAS.  相似文献   

17.
The purpose of the present study is the investigation of the structure and dynamics of biological membranes using solid-state nuclear magnetic resonance (NMR) spectroscopy. Two approaches are used in our laboratory. The first involves the measurement of high-resolution 13C and 1H spectra obtained by the magic angle spinning (MAS) technique while the second approach involves the measurement of 31P and 2H powder spectra in static samples. This paper will present some recent results obtained by high-resolution solid-state 1H NMR on the conformation of gramicidin A incorporated in a phosphatidylcholine bilayers. More specifically, we were able to observe changes in the gramicidin spectra as a function of the cosolubilization solvent initially used to prepare the samples. The interaction between lipid bilayers and an anticancer drug derived from chloroethylurea was also investigated using proton NMR spectroscopy. Finally, we have studied the interaction between cardiotoxin, a toxic protein extracted from snake venom, and negatively charged lipid bilayers using 31P solid-state NMR spectroscopy.  相似文献   

18.
The objectives of this study were to investigate the influence of physicochemical properties of lipid/plasmid complexes on in vivo gene transfer and biodistribution characteristics. Formulations based on 1,2-di-O-octadecenyl-3-trimethylammonium propane (DOTMA) and novel biodegradable cationic lipids, such as ethyl dioleoyl phosphatidylcholine (EDOPC), ethyl palmitoyl myristyl phosphatidylcholine (EPMPC), myristyl myristoyl carnitine ester (MMCE), and oleyl oleoyl L-carnitine ester (DOLCE), were assessed for gene expression after tail vein injection of lipid/plasmid complexes in mice. Gene expression was influenced by cationic lipid structure, cationic lipid-to-colipid molar ratios, plasmid-to-lipid charge ratios, and precondensation liposome size. Detectable levels of human growth hormone (hGH) in serum, human factor IX (hFIX) in plasma, and chloramphenicol acetyltransferase (CAT) in the lung and liver were observed with positively charged lipid/plasmid complexes prepared from 400-nm extruded liposomes with a cationic lipid-to-colipid ratio of 4:1 (mol/mol). Intravenous administration of lipid/CAT plasmid complexes resulted in distribution of plasmid DNA mainly to the lung at 15 min after injection. Plasmid DNA accumulation in the liver increased with time up to 24 hr postinjection. There was a 10-fold decrease in the amount of plasmid DNA in the lung at 15 min after injection, when the lipid/plasmid complex charge ratio was decreased from 3:1 to 0.5:1 (+/-). Bright fluorescent aggregates were evident in in vivo-transfected lung with the positively charged pCMV-CAT/DOLCE:dioleyl phosphatidylethanolamine (DOPE) (1:1, mol/mol) complexes, while more discrete punctate fluorescence was observed with a 4:1 molar ratio of cationic lipid:colipid formulations. Preinjection of polyanions such as plasmid, dextran sulfate, polycytidic acid, and polyinosinic acid decreased hGH expression, whereas the preinjection of both positively charged and neutral liposomes had no effect on hGH serum levels. Of the cationic lipids tested, DOLCE was found to be the most effective potentially biodegradable cationic lipid. A correlation between gene expression and cationic lipid:colipid ratios and lipid-to-plasmid charge ratio was also observed for DOTMA- and DOLCE-based formulations.  相似文献   

19.
Free thin liquid films (foam films) formed from aqueous dispersions of dimyristoylphosphatidylcholine (DMPC) and dipalmitoylphosphatidylethanolamine with covalently bound poly-(ethylene glycol) of molecular weight 2000 (DPPE-PEG-2000) were studied by the thin liquid film microinterferometric technique of Scheludko and Exerowa in the temperature range 14-36 degrees C. The surface tension kinetics of the dispersions were studied in order to ensure equilibration of the foam films. These measurements showed that the rate of surface coverage depends slightly on the temperature and does not reach equilibrium values within reasonable time intervals for the dispersions containing only one amphiphile (DPPE-PEG-2000). The destruction of the vesicles at the air/(aqueous dispersion) interface was much faster for the dispersions containing DMPC/DPPE-PEG-2000 mixtures above 23 degrees C, the temperature of the chain-melting phase transition of the main lipid component (DMPC). The dependence of the equilibrium thickness of the foam films on the electrolyte concentration was measured for 1 and 9 mol% DPPE-PEG-2000 at 28 degrees C in the range 10-3 to 0.5 M NaCl. These results indicate that at the low electrolyte concentrations the electrostatic and van der Waals interactions are dominant similar to the foam films stabilized with DMPC alone. At the high electrolyte concentrations the steric repulsion of the PEG layers becomes dominant. The temperature-composition dependence of the bilayer thickness was measured for the foam bilayers at 0.14 M NaCl. The data for the foam bilayer thickness and the comparison with the phase diagrams of PC/PE-PEG dispersions, show that the DMPC/DPPE-PEG-2000 foam bilayers are able to exist in two phase states characterised by different conformations (mushroom and brush) of the grafted polymer.  相似文献   

20.
We have investigated the influence of the chemical structure and charge of the hydrophillic headgroup on the miscibility of saturated phospholipids with acyl chain lengths differing by two methylene units, namely DMPA/DPPA, DMPC/DPPC, DMPE/DPPE and DMPG/DPPG (0.1 M NaCl). All four mixtures were analysed by DSC at pH 7. To study the influence of a change in headgroup charge, we additionally investigated DMPA/DPPA mixtures at pH 4 and 12, and DMPG/DPPG mixtures at pH 2. The experimental DSC thermograms were fitted using methods described before [Johann et al., Biophys. J. 71 (1996), 3215-3228] to obtain the temperatures of onset and end of melting and first approximations for the non-ideality parameters as a function of composition. The resulting phase diagrams were then fitted using a four non-ideality parameter model for non-ideal, non-symmetric mixing in both phases. The phase diagram of the system DMPG/DPPG has a lens-like shape, the non-ideality parameters rhog and rhol for the gel and the liquid-crystalline phase, respectively, are zero, indicating ideal mixing in both phases. For the other mixtures, differences in miscibility are observed depending on the structure of the headgroup. At pH 7, rhog > rhol, i.e., the miscibility in the liquid-crystalline phase is more ideal than in the gel state. All rhog values are positive and the sequence for rhog observed is PA>PE>PC>PG. Partial protonation of PA at pH 4 or complete deprotonation at pH 12 leads to negative non-ideality parameters for both phases, indicating a preference for mixed pair formation. Protonation of PG in DMPG/DPPG mixtures at pH 2 leads to positive non-ideality parameters for both phases, indicating a tendency for demixing. The results show, that the miscibility of phospholipids with identical headgroups but chain lengths differing by two methylene groups is dependent on headgroup structure and on headgroup charge.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号